You are on page 1of 8

Holomorphic function

In mathematics, a holomorphic function is a complex-valued function of one or more complex


variables that is, at every point of its domain, complex differentiable in a neighborhood of the
point. The existence of a complex derivative in a neighbourhood is a very strong condition, for
it implies that any holomorphic function is actually infinitely differentiable and equal, locally, to
its own Taylor series (analytic). Holomorphic functions are the central objects of study in
complex analysis.

A rectangular grid (top) and its


image under a conformal map f
(bottom).

Though the term analytic function is often used interchangeably with "holomorphic function",
the word "analytic" is defined in a broader sense to denote any function (real, complex, or of
more general type) that can be written as a convergent power series in a neighbourhood of
each point in its domain. The fact that all holomorphic functions are complex analytic
functions, and vice versa, is a major theorem in complex analysis.[1]

Holomorphic functions are also sometimes referred to as regular functions.[2] A holomorphic


function whose domain is the whole complex plane is called an entire function. The phrase
"holomorphic at a point z0" means not just differentiable at z0, but differentiable everywhere
within some neighbourhood of z0 in the complex plane.

Definition
The function is not
complex-differentiable at zero,
because as shown above, the value
of varies depending

on the direction from which zero is


approached. Along the real axis, f
equals the function g(z) = z and the
limit is 1, while along the imaginary
axis, f equals h(z) = −z and the limit
is −1. Other directions yield yet
other limits.

Given a complex-valued function f of a single complex variable, the derivative of f at a point z0


in its domain is defined by the limit[3]

This is the same as the definition of the derivative for real functions, except that all of the
quantities are complex. In particular, the limit is taken as the complex number z approaches z0,
and must have the same value for any sequence of complex values for z that approach z0 on
the complex plane. If the limit exists, we say that f is complex-differentiable at the point z0.
This concept of complex differentiability shares several properties with real differentiability: it
is linear and obeys the product rule, quotient rule, and chain rule.[4]

If f is complex differentiable at every point z0 in an open set U, we say that f is holomorphic on


U. We say that f is holomorphic at the point z0 if f is complex differentiable on some
neighbourhood of z0.[5] We say that f is holomorphic on some non-open set A if it is
holomorphic in an open set containing A. As a pathological non-example, the function given by
f(z) = |z|2 is complex differentiable at exactly one point (z0 = 0), and for this reason, it is not
holomorphic at 0 because there is no open set around 0 on which f is complex differentiable.

The relationship between real differentiability and complex differentiability is the following. If a
complex function f(x + i y) = u(x, y) + i v(x, y) is holomorphic, then u and v have first partial
derivatives with respect to x and y, and satisfy the Cauchy–Riemann equations:[6]

or, equivalently, the Wirtinger derivative of f with respect to the complex conjugate of z is
zero:[7]

which is to say that, roughly, f is functionally independent from the complex conjugate of z.

If continuity is not given, the converse is not necessarily true. A simple converse is that if u and
v have continuous first partial derivatives and satisfy the Cauchy–Riemann equations, then f is
holomorphic. A more satisfying converse, which is much harder to prove, is the Looman–
Menchoff theorem: if f is continuous, u and v have first partial derivatives (but not necessarily
continuous), and they satisfy the Cauchy–Riemann equations, then f is holomorphic.[8]

Terminology

The word "holomorphic" was introduced by two of Cauchy's students, Briot (1817–1882) and
Bouquet (1819–1895), and derives from the Greek ὅλος (holos) meaning "entire", and µορφή
(morphē) meaning "form" or "appearance".[9]

Today, the term "holomorphic function" is sometimes preferred to "analytic function". An


important result in complex analysis is that every holomorphic function is complex analytic, a
fact that does not follow obviously from the definitions. The term "analytic" is however also in
wide use.

Properties

Because complex differentiation is linear and obeys the product, quotient, and chain rules, the
sums, products and compositions of holomorphic functions are holomorphic, and the quotient
of two holomorphic functions is holomorphic wherever the denominator is not zero.[10]
If one identifies C with R2, then the holomorphic functions coincide with those functions of two
real variables with continuous first derivatives which solve the Cauchy–Riemann equations, a
set of two partial differential equations.[6]

Every holomorphic function can be separated into its real and imaginary parts, and each of
these is a solution of Laplace's equation on R2. In other words, if we express a holomorphic
function f(z) as u(x, y) + i v(x, y) both u and v are harmonic functions, where v is the harmonic
conjugate of u.[11]

Cauchy's integral theorem implies that the contour integral of every holomorphic function
along a loop vanishes:[12]

Here γ is a rectifiable path in a simply connected open subset U of the complex plane C whose
start point is equal to its end point, and f : U → C is a holomorphic function.

Cauchy's integral formula states that every function holomorphic inside a disk is completely
determined by its values on the disk's boundary.[12] Furthermore: Suppose U is an open subset
of C, f : U → C is a holomorphic function and the closed disk D = {z : |z − z0 | ≤ r} is completely
contained in U. Let γ be the circle forming the boundary of D. Then for every a in the interior of
D:

where the contour integral is taken counter-clockwise.

The derivative f′(a) can be written as a contour integral[12] using Cauchy's differentiation
formula:

for any simple loop positively winding once around a, and

for infinitesimal positive loops γ around a.


In regions where the first derivative is not zero, holomorphic functions are conformal in the
sense that they preserve angles and the shape (but not size) of small figures.[13]

Every holomorphic function is analytic. That is, a holomorphic function f has derivatives of
every order at each point a in its domain, and it coincides with its own Taylor series at a in a
neighbourhood of a. In fact, f coincides with its Taylor series at a in any disk centred at that
point and lying within the domain of the function.

From an algebraic point of view, the set of holomorphic functions on an open set is a
commutative ring and a complex vector space. Additionally, the set of holomorphic functions in
an open set U is an integral domain if and only if the open set U is connected. [7] In fact, it is a
locally convex topological vector space, with the seminorms being the suprema on compact
subsets.

From a geometric perspective, a function f is holomorphic at z0 if and only if its exterior


derivative df in a neighbourhood U of z0 is equal to f′(z) dz for some continuous function f′. It
follows from

that df′ is also proportional to dz, implying that the derivative f′ is itself holomorphic and thus
that f is infinitely differentiable. Similarly, the fact that d(f dz) = f′ dz ∧ dz = 0 implies that any
function f that is holomorphic on the simply connected region U is also integrable on U. (For a
path γ from z0 to z lying entirely in U, define

in light of the Jordan curve theorem and the generalized Stokes' theorem, Fγ(z) is independent
of the particular choice of path γ, and thus F(z) is a well-defined function on U having
F(z0) = F0 and dF = f dz.)

Examples

All polynomial functions in z with complex coefficients are holomorphic on C, and so are sine,
cosine and the exponential function. (The trigonometric functions are in fact closely related to
and can be defined via the exponential function using Euler's formula). The principal branch of
the complex logarithm function is holomorphic on the set C ∖ {z ∈ R : z ≤ 0}. The square root
function can be defined as

and is therefore holomorphic wherever the logarithm log(z) is. The function 1/z is holomorphic
on {z : z ≠ 0}.

As a consequence of the Cauchy–Riemann equations, a real-valued holomorphic function must


be constant. Therefore, the absolute value of z, the argument of z, the real part of z and the
imaginary part of z are not holomorphic. Another typical example of a continuous function
which is not holomorphic is the complex conjugate z formed by complex conjugation.

Several variables

The definition of a holomorphic function generalizes to several complex variables in a


straightforward way. Let D denote an open subset of Cn, and let f : D → C. The function f is
analytic at a point p in D if there exists an open neighbourhood of p in which f is equal to a
convergent power series in n complex variables.[14] Define f to be holomorphic if it is analytic
at each point in its domain. Osgood's lemma shows (using the multivariate Cauchy integral
formula) that, for a continuous function f, this is equivalent to f being holomorphic in each
variable separately (meaning that if any n − 1 coordinates are fixed, then the restriction of f is a
holomorphic function of the remaining coordinate). The much deeper Hartogs' theorem proves
that the continuity hypothesis is unnecessary: f is holomorphic if and only if it is holomorphic in
each variable separately.

More generally, a function of several complex variables that is square integrable over every
compact subset of its domain is analytic if and only if it satisfies the Cauchy–Riemann
equations in the sense of distributions.

Functions of several complex variables are in some basic ways more complicated than
functions of a single complex variable. For example, the region of convergence of a power
series is not necessarily an open ball; these regions are Reinhardt domains, the simplest
example of which is a polydisk. However, they also come with some fundamental restrictions.
Unlike functions of a single complex variable, the possible domains on which there are
holomorphic functions that cannot be extended to larger domains are highly limited. Such a set
is called a domain of holomorphy.

A complex differential (p,0)-form α is holomorphic if and only if its antiholomorphic Dolbeault


derivative is zero, .

Extension to functional analysis

The concept of a holomorphic function can be extended to the infinite-dimensional spaces of


functional analysis. For instance, the Fréchet or Gateaux derivative can be used to define a
notion of a holomorphic function on a Banach space over the field of complex numbers.
See also

Antiderivative (complex analysis)

Antiholomorphic function

Biholomorphy

Holomorphic separability

Meromorphic function

Quadrature domains

Harmonic maps

Harmonic morphisms

Wirtinger derivatives

References

1. Analytic functions of one complex variable , Encyclopedia of Mathematics. (European


Mathematical Society ft. Springer, 2015)

2. Springer Online Reference Books , Wolfram MathWorld

3. Ahlfors, L., Complex Analysis, 3 ed. (McGraw-Hill, 1979).

4. Henrici, P., Applied and Computational Complex Analysis (Wiley). [Three volumes: 1974,
1977, 1986.]

5. Peter Ebenfelt, Norbert Hungerbühler, Joseph J. Kohn, Ngaiming Mok, Emil J. Straube
(2011) Complex Analysis Springer Science & Business Media

6. Markushevich, A.I.,Theory of Functions of a Complex Variable (Prentice-Hall, 1965).


[Three volumes.]

7. Gunning, Robert C.; Rossi, Hugo (1965), Analytic Functions of Several Complex
Variables , Prentice-Hall series in Modern Analysis, Englewood Cliffs, N.J.: Prentice-Hall,
pp. xiv+317, ISBN 9780821869536, MR 0180696 , Zbl 0141.08601

8. Gray, J. D.; Morris, S. A. (1978), "When is a Function that Satisfies the Cauchy-Riemann
Equations Analytic?", The American Mathematical Monthly (published April 1978), 85 (4):
246–256, doi:10.2307/2321164 , JSTOR 2321164 .

9. Markushevich, A. I. (2005) [1977]. Silverman, Richard A. (ed.). Theory of functions of a


Complex Variable (2nd ed.). New York: American Mathematical Society. p. 112. ISBN 0-
8218-3780-X.
10. Henrici, Peter (1993) [1986], Applied and Computational Complex Analysis Volume 3 ,
Wiley Classics Library (Reprint ed.), New York - Chichester - Brisbane - Toronto -
Singapore: John Wiley & Sons, pp. X+637, ISBN 0-471-58986-1, MR 0822470 ,
Zbl 1107.30300 .

11. Evans, Lawrence C. (1998), Partial Differential Equations, American Mathematical Society.

12. Lang, Serge (2003), Complex Analysis, Springer Verlag GTM, Springer Verlag

13. Rudin, Walter (1987), Real and complex analysis (3rd ed.), New York: McGraw–Hill Book
Co., ISBN 978-0-07-054234-1, MR 0924157

14. Gunning and Rossi, Analytic Functions of Several Complex Variables, p. 2.

Further reading

Blakey, Joseph (1958). University Mathematics (2nd ed.). London: Blackie and Sons.
OCLC 2370110 .

External links

"Analytic function" , Encyclopedia of Mathematics, EMS Press, 2001 [1994]

You might also like